Вы находитесь на странице: 1из 7

High-temperature deformation and fracture of B i - S r - C a - C u - 0 superconductors

K. C. Goretta and E. J. Zamirowski


Energy Technology Division, Argonne National Laboratory, Argonne, Illinois 60439-4838

J. M. Calderon-Moreno
Departamento de Fisica de la Materia Condensada, Universidad de Sevilla, 41080 Sevilla, Spain

D. J. Miller, Nan Chen, T. G. Holesinger, and J. L. Routbort


Materials Science Division, Argonne National Laboratory, Argonne, Illinois 60439-4838 (Received 28 May 1993; accepted 16 October 1993)

Dense polycrystalline Bi 2 Sr 2 Cu0 x (2201), Bi 2 Sr 2 CaCu 2 0 x (2212), and (Bi,Pb) 2 Sr 2 Ca 2 Cu 3 O x (2223) specimens were compressed in air at 730-835 C. All of the materials exhibited an apparent steady-state creep response. Strain rate was proportional to stress to the 3.1-3.8 power. Apparent activation energies for the deformation processes were 520 50 kJ/mole for the 2201, 630 210 kJ/mole for the 2212, and 960 2 1 0 kJ/mole for the 2223. Transmission electron microscopy revealed substantial generation and propagation of basal-plane dislocations during deformation. Few nonbasal-plane dislocations were observed. Intergranular fracture was evident in all deformed samples, and intragranular fracture was evident along the basal planes of some grains. It is suggested that the kinetics of fracture were determined by dislocation motion within the grains.

I. INTRODUCTION High-temperature deformation has been used to densify polycrystalline Bi-based superconductors.1"7 Improvements in transport critical current density, Jc, for these conductors compared with conventionally sintered products have been reported. 26 In addition, dislocation generation and motion occur during deformation.7 It has been shown that large concentrations of dislocations can increase intragranular Jc in Bi-based superconductors. 810 Therefore, it is possible that hightemperature deformation can improve both grain connectivity and intragranular properties as compared with conventional sintering. Previous studies of compressive creep of Bi 2Sr 2CaCu 2 O x have indicated that deformation can occur by dislocation motion.11"13 Microstructural examinations have revealed, however, that cracking generally accompanies the deformation. In addition, it has been shown that Bi 2 Sr 2 CaCu 2 0 x can deform by diffusional creep without cracking for stresses lower than 10 MPa.13 Of the superconducting phases in the B i - S r - C a C u - 0 system, the nominal compositions Bi 2 Sr 2 Cu0^ (2201), Bi 2 Sr 2 CaCu 2 0^ (2212), and (Bi, Pb) 2 Sr 2 Ca 2 C\1T,OX (2223) have been synthesized in a highly pure form and have been extensively studied. Phases with more C u - 0 layers have been identified and made, but it remains difficult to obtain them in pure form.14 2201, 2212, and 2223 differ structurally from each other primarily by the number of superconducting C u - O layers they contain. The goals of this study were to

determine the extents to which Bi-based superconductors could be deformed by dislocation motion and to which they deformed similarly, the kinetics of the deformation processes, and whether the plastic deformation was accompanied by fracture. To accomplish these goals, dense specimens were fabricated from 2201, 2212, and 2223 powders. Compressive creep tests were performed in air over similar ranges of temperature and compression rate, and the data were supplemented by electron microscopy observations to determine the nature of the dislocations and fractures.

II. EXPERIMENTAL DETAILS The 2201, 2212, and 2223 powders were synthesized by a reaction of mixtures of appropriate amounts of Bi 2 O 3 , PbO, SrCO 3 , CaCO 3 , and CuO. The actual compositions were Bi^nSr^gCuO*, Bi2Sr225Cao.9Cu2Ox, and Bij 8 Pb 0 .4Sr 2 Ca 22 Cu 3 O^, which yielded powders of good phase purity.15"21 The powders were ball-milled for 16-24 h in polyethylene jars containing methyl or isopropyl alcohol and ZrO 2 grinding media. The milled powders were pan-dried, ground lightly with an agate mortar and pestle, placed into A12O3 crucibles, and heated by various schedules. The 2201 was heated in air at 735 C for 23 h and then at 805 C for 40 h. The 2212 was prepared by calcining the powder mixture in air, and then melting the mixture in an A12O3 crucible, followed by splat quenching. The 2223 was heated in air at 805 C for 40 h and then in 8% O 2 /92% Ar at
541

J. Mater. Res., Vol. 9, No. 3, Mar 1994

K. C. Goretta et al.: High-temperature deformation and fracture of Bi-Sr-Ca-Cu-0 superconductors

835 C for 50 h. X-ray diffraction indicated only trace impurities in each of the powders.21 To ensure that all deformation tests were performed in the solid state, differential thermal analysis (DTA) was used to examine the melting behaviors of the powders, and the melting temperatures were not exceeded during deformation. The 2201 powder was processed into bulk form by hot isostatic pressing (HIPing). The HIPing was conducted for 2 h in an inert atmosphere at 105 MPa and 825 C. n Large 34 mm thick disks were processed; several test specimens were cut from each with a slowspeed diamond saw. The specimens were fully dense and, as revealed by x-ray diffraction, only modestly textured.22 Very little second phase was observed by scanning and transmission electron microscopy (SEM and TEM). The 2212 specimens were prepared by a directionalsolidification technique.23 The splat-quenched material was ground, placed in an Ag-lined A12O3 crucible, melted at 900 C, cooled to 850 C, and held for 100 h in flowing O 2 . The specimens were 92% dense and highly textured.22 SEM of the cast sample revealed a modest amount of SrO and (Sr, C a ^ C u ^ O ^ , and a trace of Al-containing phases that arose from Al incorporation during initial melting in the A12O3 crucible.24 The 2223 was sinter forged in air at 850 C.22'25 The compression rates were 0.005-0.10 mm/min; the final stress was =3 MPa. The products were 90-95% dense and very highly textured.22'26 SEM revealed that the specimens consisted of elongated 2223 grains and a few percent elongated 2212 grains and blocky alkalineearth cuprates. The deformation specimens were cut from the bar by a slow-speed diamond saw; subsequent compression occurred parallel to the direction of elongation of the grains. Test specimens were approximately 2 X 2 mm or 3 X 3 mm in cross section and were 1.5 times taller than wide. They were compressed in air between A12O3 platens at a constant crosshead velocity in a testing machine equipped with a high-temperature furnace and a muffle tube.27-28 Samples were equilibrated for ~ 1 h at the test temperature prior to compression. Test temperatures were 730-835 C. Initial strain rates ranged from 2 X 10~6 to 2 X 10" 5 s" 1 . The specimens were deformed until an apparent steady state was reached. Steady-state stress, cr, or strain rate, e, was defined by a zero work-hardening rate.29 The apparent steady states were achieved within 2% axial strain. For most specimens, e or temperature was changed during the test and more than one apparent steady state was established. Slight barrelling occurred during compression as the result of friction. No barrelling correction was made; all strain rates were calculated from the initial sample lengths. The increments of strain between individual tests were always less than
542

2%, and thus no significant error was introduced by the use of approximate stresses and strains.27 Samples for TEM were prepared by slicing with a slow-speed diamond saw. Deformed samples were cut at an angle of ~20 to the compression axis. The samples were thinned, dimpled, and polished. The thinned samples were ion milled to final perforation with a 4 kV Ar beam at liquid-N2 temperature. TEM was performed on a Philips CM-30 operated at 100 kV. III. RESULTS AND DISCUSSION Data from a representative 2201 test are shown in Fig. 1. Within experimental error, stress (cr) was approximately proportional to strain rate (e). The data shown are representative of all specimens; apparent steady states were always obtained. In general, however, total strains for the specimens were limited to = 5 % . The data were fit to the standard creep equation, = A^ex^-Q/RT),
(1)

where A is a constant, n the stress exponent, Q the activation energy, R the gas constant, and T the absolute temperature.29 The apparent steady-state stresses plotted versus strain rate are shown for 2201 in Fig. 2, for 2212 in Fig. 3, and for 2223 in Fig. 4. The calculated values of n are shown on the plots. The 2201 data exhibited a range of scatter typical for creep tests of superconductors.26"28 The calculated value of n was within an experimental error of 3.0, which is common for a dislocation-creep mechanism.29 The 2212 data exhibited substantially more scatter. For example, the two specimens deformed at 780 C were stronger

HI 4 ca 2=
II

II

i i
. . . . i

2 "

1 "

0.02

0.04

0.06

0.08

0.1

Strain
FIG. 1. Stress-strain data for 2201 samples compressed at 811 C; the strain rates for the various tests were II = 1 X 10"5 s"

J. Mater. Res., Vol. 9, No. 3, Mar 1994

K. C. Goretta et al.: High-temperature deformation and fracture of B i - S r - C a - C u - 0 superconductors

1.2
758C
1.0

2.0

730C

1.5

o o o

-o
779C

0.8 "

1 *
A
A A

*
A 784C

A &
A
m

8 rc 1 A
834C

1.0

A
A

803C

35

0.6

' A
A

81
0.5

n = 3.8 0.7

n = 3.1 0.3

0.4

-4.2 0.0 -5.8 -5.6 -5.4 -5.2

Z1 n = 2.4

827C

0.2 -5.4

-5.2

-5.0

-4.8

-4.6

-4.4

-5.0

-4.8

-4.6

log (Strain Rate [/s])

log (Strain Rate [/s])

FIG. 2. Apparent steady-state stress versus strain rate for 2201 specimens compressed at the temperatures shown.

FIG. 4. Apparent steady-state stress versus strain rate for 2223 specimens compressed at the temperatures shown.

than the one deformed at 760 CC. The error bars on the value of n reflect the inherent scatter in the data. Within experimental error, however, n was approximately 3. The 2223 data were more consistent than the 2212 data, but the specimen deformed at the highest temperature was much softer than the other specimens and its n value was much lower. The scatter in the data is probably related to the relatively low densities and phase purities of the 2212 and 2223 specimens relative to those of the 2201 specimens. It is believed that the lower n for the 2223 at the highest temperature was caused by a change in the mechanism of deformation. 2223 has the lowest melting temperature of the three compounds. In air,
2.0 753C

1.8

A A

780C 760C 790C

9*
A
804C

co

1.6 "

the melting points determined by DTA were 2201 ~ 895 C, 2212 = 885 C, and 2223 865 C. However, DTA is not very effective in detecting small amounts of liquid.30 In addition, the specimens were prepared in Ag foils and the presence of Ag can lower the melting points of B i - S r - C a - C u - 0 superconductors by ~20 C.31'32 A small amount of melting may have caused a reduction in the value of n for 2223 at 827 C. It is also possible that diffusional flow, for which n typically equals one, may have become appreciable at that temperature,14 and that n = 2.4 resulted from a mixture of n = 1 and n = 3 mechanisms. In any case, the data for that one test were sufficiently outside the trends for the other tests that they were omitted from calculations for apparent activation energies. Apparent activation energies for the deformation processes were calculated from the slopes of plots of n In cr versus l/T. The values are shown in Table I. The error bars are due largely to uncertainties in the value of n. The apparent Q values of each compound show excellent agreement for various strain rates. A
TABLE I. Apparent activation energies in air for high-temperature deformation of Bi-Sr-Ca-Cu-O superconductors obtained for various compressive strain rates.

44440

CO,

0)

A
1.4

A
1.2
n = 3.6 0.9 1.0 -5.8

823C

Specimen 2201 2201 2201 2212 2212 2223 2223

e(s-l)
5 1 2 6 1 5 1 X X X X X X X 10"6 10"5 10"5 10"6 10^5 10~6 10~5

Q (kJ/mole)
520 510 530 625 640 980 950 50 50 50 200 200 200 200

-5.6

-5.4

-5.2

-5.0

-4.8

-4.6

log (Strain Rate [/s])

FIG. 3. Apparent steady-state stress versus strain rate for 2212 specimens compressed at the temperatures shown.

J. Mater. Res., Vol. 9, No. 3, Mar 1994

543

K. C. Goretta et al.: High-temperature deformation and fracture of B i - S r - C a - C u - 0 superconductors

10 Stress (MPa)

100

FIG. 5. Strain rate versus stress plot for HIPed 2212 deformed at lower stresses (n ~ 1 regime) and melt-cast 2212 deformed at higher stresses (n = 3 regime) at the temperatures shown.

trend of increasing Q with the number of C u - 0 layers (and hence decreased average interplanar spacing) seems clear. However, examination of Figs. 2 - 4 reveals that the 2201 was the softest and that, despite its intermediate apparent activation energy, the 2212 was the hardest. Although it is possible that the relative strengths were a consequence of extrinsic effects, comparison of the data presented here with those from our previous study14 militate against that possibility. In the previous study, HIPed

2212 polycrystals of composition Bi2Sr1.7CaCu2OJ: were compressed at low stresses and deformed by diffusional creep. The 2212 specimens used here were melt cast and had a different composition. Yet, as shown in Fig. 5, the two sets of data agree very well, with the higher stresses inducing a higher stress exponent. Effects of texture might have had an effect on the apparent steady-state stresses. However, when the 2212 was compressed in two orientations, one along the preferred direction of the c axes and one perpendicular to the c axes, the data agreed within =10%. Although the data for the 2212 and 2223 specimens exhibited significant scatter, and the flow stresses were not proportional to Q, some form of steady-state appears to have been established. In all deformed specimens, TEM revealed cracks. Many of the cracks were intergranular, but some were intragranular. In contrast, no intragranular cracks were observed in the as-HIPed specimens, although some intergranular cracks were seen in those samples. The intragranular cracks in the deformed samples were most commonly observed to lie along directions in the a-b planes, as seen in Fig. 6. Many of the grains in the deformed samples contained high concentrations of dislocations (Fig. 7), but the distribution of dislocations was not uniform and some grains contained very few. The undeformed specimens contained relatively low concentrations of dislocations (Fig. 8), although dislocation

FIG. 6. Bright-field TEM micrograph of a 2212 specimen deformed at 750 C, imaged with g = (002), as indicated by the arrow. Several intragranular cracks which lie between a-b planes are visible. 544 J. Mater. Res., Vol. 9, No. 3, Mar 1994

K. C. Goretta et al.: High-temperature deformation and fracture of B i - S r - C a - C u - 0 superconductors

FIG. 7. Bright-field TEM micrograph of dislocations observed in (a) 2201 compressed at 785 C, with the central grain imaged with g = (200), as indicated; and (b) 2212 compressed at 830 C, with the grain to the left imaged with g = (115), as indicated.

networks were occasionally observed at grain boundaries. These networks were typically associated with [001] twist boundaries.33 In addition, a few dislocations could be found within grains. These dislocations were generally long and straight and were likely to have been introduced during the HIPing process.

The dislocations and faults observed in the deformed samples were primarily, but not entirely, confined to the basal planes. These dislocations are characteristic of those formed by glide rather than by climb. In Fig. 7(b), for example, many of the dislocations appear to have advanced by a kink mechanism, leaving behind a stack545

J. Mater. Res., Vol. 9, No. 3, Mar 1994

K. C. Goretta et al.: High-temperature deformation and fracture of B i - S r - C a - C u - 0 superconductors

200 nm
FIG. 8. Bright-field TEM micrograph of an undeformed 2223 specimen, with the central grain imaged with g = (020), as indicated.

ing fault. In some places, jog-type motion was evident and preliminary analyses indicate that a few dislocations contain a Burgers vector with a c-axis component. For example, the image contrast for the faults and dislocations in Fig. 6 is due to a nonbasal-plane component for which g b or g R is finite with g = (002). Weak-beam imaging has shown narrow dislocation images bounding these faults, suggesting that these faults may be extended by glide of these dislocations. The dislocations that were observed here are similar to those observed previously in B i - S r - C a - C u - 0 superconductors in that most of them were confined to the basal plane.34"38 Further work is underway to fully characterize the nature of the observed defects, especially the nonbasal components to the dislocations. The contribution of all of the dislocations to the total strain was insufficient to allow for general plasticity. Although some strain was accommodated by dislocation glide, microcracks were introduced to relieve the applied stress. Microcracking occurred primarily by cleavage between basal planes, but was not limited to that orientation. The question remains as to why specimens deforming by dislocation glide and fracture exhibited responses that are characteristic of steady-state creep. It is possible that the kinetics of the sequential microcracking were governed by bending of the individual grains and that the bending of the grains was controlled by dislocation glide. The increase in apparent activation energy for the deformation process with increasing complexity of the
546

crystal structure is consistent with this argument. With regard to the creep parameters, n values are typical of a dislocation mechanism29 and the relatively high values of Q are consistent with creep of high-temperature superconductors.12"14'26"28'39'40 The scatter in the apparent steady-state-stress data is probably the result of differences in cracking. The 2201 had the most homogeneous microstructure and hence probably microcracked most uniformly. Because the dislocations were largely confined to the basal planes and high concentrations of dislocations were associated with fracture, it appears to be unlikely that creation of dislocations by uniaxial pressure can be used to improve Jc of B i - S r - C a - C u - 0 superconductors. Isostatic deformation would probably be more effective, although its benefits also seem to be limited.11 IV. CONCLUSIONS Dense 2201, 2212, and 2223 specimens were compressed in air from 730-835 C. All of the materials appeared to deform by steady-state creep, with strain rates proportional to stress to the 3.1-3.8 power and apparent activation energies of 520 50 kJ/mole for the 2201, 630 210 kJ/mole for the 2212, and 960 210 kJ/mole for the 2223. TEM revealed the simultaneous occurrence of dislocation glide and microcracking. It is suggested that the kinetics of fracture were determined by bending of the grains caused by dislocation motion.

J. Mater. Res., Vol. 9, No. 3, Mar 1994

K. C. Goretta et a/.: High-temperature deformation and fracture of B i - S r - C a - C u - 0 superconductors

ACKNOWLEDGMENTS We thank S.E. Dorris for providing the 2223 powder and L. J. Martin for preparing some of the specimens. This work was supported by the United States Department of Energy, Energy Efficiency and Renewable Energy, as part of a Department of Energy program to develop electric power technology, and Basic Energy Sciences-Materials Sciences, under Contract W31-109-ENG-38. E.J.Z. was partially supported by the Argonne Division of Educational Programs through funding from the Department of Energy. J. M. C.-M. was supported by Spanish Grant CICYT MAT88-0181-C02 (Ministerio de Education y Ciencias). REFERENCES
1. N. Murayama and Y. Torii, Jpn. J. Appl. Phys. 28, L1763 (1989). 2. G. A. Whitlow, W. R. Lovic, and J. C. Bowker, Supercond. Sci. Technol. 4, 353 (1991). 3. A. Tampieri and G. N. Babini, Jpn. J. Appl. Phys. 30, LI 163 (1991). 4. K. M. Demchuk, G. K. Pokazan'eva, S.A. Lebedev, S.M. Cheshnitskii, A. N. Martem'yanov, B.I. Kamenetskii, V.I. Levit, L.S. Davydova, M.I. Tsypin, and L.T. Kir'yanova, Superconductivity 4, 1883 (1991). 5. B.M. Moon, V.N. Korenivski, K.V. Rao, and Y. J. Kim, Mater. Lett. 14, 72 (1992). 6. X. Wu and I-W. Chen, J. Am. Ceram. Soc. 75, 1846 (1992). 7. K. C. Goretta, D.J. Miller, R. B. Poeppel, and A. S. Nash, in Pressure Effects on Materials Processing and Design, edited by K. Ishizaki, E. Hodge, and M. Concannon (Mater. Res. Soc. Symp. Proc. 251, Pittsburgh, PA, 1992), p. 325. 8. S. X. Dou, H. K. Liu, J. Wang, and W. M. Bian, Supercond. Sci. Technol. 4, 21 (1991). 9. H. K. Liu, Y. C. Guo, and S. X. Dou, Supercond. Sci. Technol. 5, 591 (1992). 10. D. J. Miller, S. Sengupta, D. Shi, J. D. Hettinger, K. E. Gray, A. S. Nash, and K. C. Goretta, Appl. Phys. Lett. 61, 2823 (1992). 11. P.E. Reyes-Morel, X. Wu, and I-W. Chen, in Ceramic Superconductors II, edited by M. F. Yan (The American Ceramics Society, Westerville, OH, 1988), p. 590. 12. V.N. Kuznetsov, V.I. Shalaev, V.V. Sagaradze, E.V. Dus'e, V. L. Arbuzov, A. E. Davletshin, V. R. Paskrebysheva, and S. M. Cheshnitskii, Superconductivity 5, 1850 (1992). 13. J.L. Routbort, K.C. Goretta, D.J. Miller, D.B. Kazelas, C. Clauss, and A. Dominguez-Rodriguez, J. Mater. Res. 7, 2360 (1992). 14. H. Wang, X. Wang, S. Shang, Z. Wang, Z. Lu, and M. Jiang, Appl. Phys. Lett. 57, 710 (1990). 15. A. K. Cheetham, A. M. Chippendale, and S. J. Hibble, Nature 333, 21 (1988).

16. S. Koyama, U. Endo, and T. Kawai, Jpn. J. Appl. Phys. 27, L1861 (1988). 17. B. Hong, J. Hahn, and T. O. Mason, J. Am. Ceram. Soc. 73, 1965 (1990). 18. B. Hong and T. O. Mason, J. Am. Ceram. Soc. 74, 1045 (1991). 19. Z. Hiroi, Y. Ikeda, M. Takano, and Y. Bando, J. Mater. Res. 6, 435 (1991). 20. T. G. Holesinger, D. J. Miller, L. S. Chumbley, M. J. Kramer, and K. W. Dennis, Physica C 202, 109 (1992). 21. D.Y. Kaufman, M.T. Lanagan, S.E. Dorris, J.T. Dawley, I.D. Bloom, M. C. Hash, N. Chen, M. R. DeGuire, and R. B. Poeppel, Appl. Supercond. 1, 81 (1993). 22. C-Y. Chu, J. L. Routbort, N. Chen, A. C. Biondo, D. S. Kupperman, and K.C. Goretta, Supercond. Sci. Technol. 5, 306 (1992). 23. T.G. Holesinger, D.J. Miller, H.K. Viswanathan, and L.S. Chumbley, J. Mater. Res. 8, 2149 (1993). 24. T. G. Holesinger, D. J. Miller, S. Fleshier, and L. S. Chumbley, J. Mater. Res. 7, 2035 (1992). 25. K. Matsuzaki, A. Inoue, and T. Masumoto, Jpn. J. Appl. Phys. 29, L1789 (1990). 26. N. Chen, Argonne National Laboratory, unpublished results (1993). 27. K. C. Goretta, J. L Routbort, A. C. Biondo, Y. Gao, A. R. de Arellano-Lopez, and A. Dominguez-Rodriguez, J. Mater. Res. 5, 2766 (1990). 28. N. Chen, S.J. Rothman, J.L. Routbort, and K.C. Goretta, J. Mater. Res. 7, 2308 (1992). 29. W.R. Cannon and T.G. Langdon, J. Mater. Sci. 18, 1 (1983). 30. P. E. D. Morgan, M. Okada, T. Matsumoto, and A. Soeta, in Advances in Superconductivity II, edited by T. Ishiguro and K. Kajimura (Springer-Verlag, Tokyo, 1990), p. 435. 31. S.X. Dou, K.H. Song, H.K. Liu, C.C. Sorrell, M.H. Apperley, A. J. Gouch, N. Savvides, and D. W. Hensley, Physica C 160, 533 (1989). 32. A. S. Nash, P. Nash, R. B. Poeppel, and K. C. Goretta, in High Temperature Superconducting Compounds III, edited by S. H. Whang, A. DasGupta, and E. W. Collings (The Minerals, Metals, and Materials Society, Warrendale, PA, 1991), p. 77. 33. Y. Gao, C-T. Wu, Y. Shi, K. L. Merkle, and K. C. Goretta, Appl. Supercond. 1, 131 (1993). 34. Z-Q. Li, H. Shen, Y. Qin, J-Y. Jiang, and J-J. Du, Philos. Mag. Lett. 60, 123 (1989). 35. C. Song, F. Liu, H. Gu, T. Lin, J. Zhang, G. Xiong, and D. Yin, J. Mater. Sci. 26, 11 (1991). 36. P. Shang, G. Yang, I. P. Jones, C. E. Gough, and J. S. Abell, Appl. Phys. Lett. 63, 827 (1993). 37. T. Murase, K. Kuroda, K. Suzuki, and H. Saka, Philos. Mag. A 62, 583 (1990). 38. Y. Takahashi and T. Suga, Jpn. J. Appl. Phys. 29, L2006 (1990). 39. A. de Arellano-Lopez, K. C. Goretta, J. L. Routbort, D. J. Miller, and A. Dominguez-Rodriguez, Ceram. Acta 3, 5 (1991). 40. J. L. Routbort, D. J. Miller, E. J. Zamirowski, and K. C. Goretta, Supercond. Sci. Technol. 6, 337 (1993).

J. Mater. Res., Vol. 9, No. 3, Mar 1994

547

Вам также может понравиться